You are on page 1of 14

III: CENTRAL FORCES and ORBITS

1st year CLASSICAL MECHANICS

HT09 RCED

A central force is one in which the force depends only on the distance from the force centre
to the particle or on the distance between two interacting particles. Both Newtonian
gravitation and the Coulomb force are examples of central forces. Polar coordinates are
the natural choice for setting up the equations of motion. Although many of the systems
of interest are two-body, most of the development will assume a fixed force centre. The
extension to the full system may be done using the reduced mass.

General Remarks

Consider a particle of mass m at a vector position r from a fixed centre of force at O


(Fig. 1(a)). The equation of motion is
mr = F(r) = f (r)r.

(r is the unit vector along OP)

(1)

The characteristics of a central force are that it is directed along the radius vector and
that it depends on r = |r| not r, hence the scalar function f (r). The is for later
convenience.

Figure 1: (a) A central force; (b) motion confined to a plane.


Take the vector cross-product of Eqn 1 with r
r r =

f
r r = 0,
m

which may be integrated once wrt time to give


r r = n

where n is a constant vector.

This equation is the vector equation of a plane and it shows that the motion of the particle
is confined to a plane with normal along n (Fig. 1(b)). Mutiplying the above equation by
m gives r (mr) = L where L is the angular momentum of the particle about O and is
a constant of the motion.
Take the dot product of Eqn 1 with r to give
mr r = f (r)r r.
1



d 1 2
mr , the KE of the particle. The RHS contains r r which is the
The LHS is
dt 2
component of the velocity along the direction of r or dr/dt. Thus the equation may be
integrated wrt time to give
Z
1 2
mr = f (r)dr + constant
2
1 2
or
mr + V (r) = Etot
2
Z
Z
This is the conservation of energy with V (r) = F dr =
f (r)dr the potential
energy function and the constant, Etot , the total energy. The potential V (r) is clearly a
function of position only, so central forces are conservative.
Two constants of the motion have been identified:
the total energy of the particle, Etot ;
the angular momentum of the particle about the centre of force, L.
These two will provide powerful tools in the analysis of central force problems.

Polar Coordinates

Since central force motion is confined to a plane, the most convenient coordinate system
is plane polar coordinates with origin at the centre of force (Fig. 2(a)). The relationships

Figure 2: (a) Polar coordinates; (b) infinitesimal changes to unit vectors r , .


between polar and rectangular coordinates are
p
r = x2 + y 2
= tan1 (y/x).

x = r cos
y = r sin

For rectangular coordinates the unit vectors i, j are in fixed directions. For polar co perpendicular to it. Their
ordinates the unit vectors are: r along the radius vector;
directions change with the motion. To work out the rates of change either use geometrical
methods as shown in Fig. 2(b) or brute force starting from
r = cos i + sin j
= sin i + cos j

For example
,
r = [ sin i + cos j] =

from which

dr
= .
dt

Similarly
= r .
dt
Using these results it is now straightforward to work out the expressions for velocity
and acceleration in 2-D polar coordinates by differentiating r = rr:

r = r r + r
.

r = (
r r 2 ) r + (r + 2r )

(2)
(3)

Equations of motion

Using the results of Eqn 3 in Eqn 1 gives


m(
r r 2 ) = f (r)
= 0
m(r + 2r )

(4)
(5)

for the equations of motion of a central force in polar coordinates. The 2nd equation may
also be written as
d  2 
r = 0 giving r 2 = h,
(6)
dt
where h is a constant. This is another formulation of the conservation of angular momentum in central force motion, which allows one to deduce the 2nd of Keplers Laws of
planetary motion
KII The radius vector from the Sun sweeps out equal areas in equal times.
1
Refer to Fig. 3, in time t the radius vector r will sweep out an area A r r
2

( = t).
So
1
h
dA
= r 2 = , which is constant.
dt
2
2

Figure 3: The area swept out by the radius vector r in time t.


The other two Laws of Kepler are
KI Each planet moves in an ellipse with the Sun at one focus.
KIII The square of the period of revolution about the Sun is proportional to the cube of
the major axis of the orbit.
These follow from Newtons Laws of motion together with his inverse square law of universal gravitation, as will be demonstrated.
3

3.1

Radial equation (not in the syllabus)

The conservation of angular momentum (Eqn 6) may be used to eliminate the dependence
on in Eqn 4, giving
f (r)
h2
.
(7)
r 3 =
r
m
This equation now depends only on r note how the use of polar coordinates has enabled
one to exploit the conservation of AM in a simple and elegant way.
Although you will not be expected to reproduce the following in Prelims, the above
equation can be transformed into an equation from which the trajectory r() may be
determined. The trick is to change variables from r to u = 1/r and use
r =

1
1 du
du
u = 2 = h
2
u
u d
d

It then follows that r = h

since = hu2

2
d2 u
2 2d u

=
h
u
and Eqn 7 becomes
d2
d2

d2 u
f (u)
+u=
.
2
d
mh2 u2

(8)

This is now a standard 2nd order linear differential equation for which the complementary
function (solution to the equation with f (u) = 0) is
u() = A cos + B sin
(SHM!), with constants A, B fixed by the initial conditions. The particular integral depends on the exact form of f . For the inverse square force f (u) = u2 where is a
constant, Eqn 8 is

d2 u
+u=
.
2
d
mh2
with PI u = /(mh2 ). Thus the complete solution for the trajectory in r space is
1

=
+ A cos + B sin
r
mh2

1
=
+ A cos( 0 )
or
r
mh2

or
= 1 + e cos( 0 ),
r

(9)

where A , 0 , e are constants and = mh2 /. By a suitable choice of axes one may choose
0 = 0. The last form (Eqn 9) is a standard representation and is the polar form of an
ellipse (0 < e < 1) or hyperbola (e > 1) with coordinate origin at a focus.

Effective potential and energy conservation

Although one could pursue the above geometrical path further, it is more instructive to
return to the radial equation and see how energy conservation emerges. Multiply Eqn 7
by m and write J = |L| = mh = mr 2 for the (conserved) AM
m
r

J2
= f (r)
mr 3
4

Veff

E tot

r1

r0

r2

Figure 4: Vef f (r) for the 2-D SHM. The position of the minimum energy is at r0 and
r1 , r2 show the extremes of the orbit with energy Etot .
multiply this by r and integrate once wrt time to give
Z
1 2
J2
mr +
= f (r) dr + constant.
2
2mr 2
Z
The constant is the total energy Etot and V (r) = f (r) dr is the potential energy function
already introduced.

J2
1 2
mr +
+ V (r) = Etot .
2
2mr 2
Next introduce the effective potential Vef f
J2
1
to give mr 2 + Vef f (r) = Etot ,
(10)
2
2mr
2
which is the energy equation for a 1-D problem. The nature of the motion will depend
on the interplay between the two components that make up Vef f . At this point it will be
useful to revise the qualitative discussion on potential functions from the MT lectures.
Vef f (r) V (r) +

4.1

2-D SHM

The two dimensional harmonic oscillator with restoring force F = kr is a simple example of a central force. Following the above procedure, the first integral gives energy
conservation
1 2
J2
1
mr +
+ kr 2 = Etot
2
2
2mr
2
1 2
The potential energy function is V (r) = kr and the effective potential
2
1 2
J2
+
kr .
Vef f (r) =
2mr 2 2
5

Figure 5: Elliptical orbits: (a) for 2-D SHM with force centre at centre; (b) inverse square
force with force centre at a focus.
Fig. 4 shows Vef f (r) and its components as functions of r. Note how the angular momentum contribution drops as r 2 , while the potential energy increases as r 2p
. This means
2
1/4
that Vef f has a minimum, which is at r0 = [J /mk] with value Emin = kJ 2 /m. For
the 1-D oscillator, the particle has zero KE when it is at maximum amplitude, here only
the radial component of the velocity is zero at the extremes of the motion. There are two
extreme positions and they are shown as r1 , r2 on the diagram. As the equation of motion
can be separated in rectangular coordinates, it is not difficult to show that the orbit is an
ellipse [try it!] with orgin at the centre and r1 , r2 as the radii of the minor and major axes
respectively (see Fig. 5(a)). Note the special case with energy Emin and a fixed value of
r = r0 , the orbit is a circle of radius r0 .

4.2

Inverse square

The force is given by F = 2 r, where is a constant (> 0, attractive; < 0 repulsive).


r
The energy equation is
1 2
J2

mr +

= Etot
2
2mr 2
r
The potential energy function is V (r) = /r and the effective potential
Vef f (r) =

J2
.
2
2mr
r

If the force is repulsive, e.g. two like charges with = |q1 q2 |/(40 ), then Vef f is a
decreasing positive function of r and there are no bound states. If the force is attractive
unlike charges = +|q1 q2 |/(40 ) or Newtonian gravity = GmM then as Fig. 6
shows, the two components of Vef f (r) have opposite signs giving the possibility of bound
state solutions. The angular momentum component is positive and decreases as r 2 ,
but now the potential energy is negative with magnitude decreasing as r 1 . Vef f has a
J2

minimum at r0 =
with value Emin =
. The orbit corresponding to the minimum
m
2r0
is again a circle and for gravity Emin = GMm/(2r0 ). For Emin < Etot < 0 there are
bound elliptical orbits, but now r1 , r2 represent the distances along the major axis from
the focus to the extremes of the orbit (as shown in Fig. 5(b)). A new feature appears for
E > 0, even though the force is attractive, the motion is unbound and corresponds to
hyperbolic motion. Now the angular momentum component is always larger in magnitude
than that from the attractive force.
6

Veff

r1

r2

r0

Figure 6: Vef f (r) for the inverse square gravitational force. The position of the minimum
energy is at r0 and r1 , r2 show the extremes of a closed orbit with energy Etot < 0.

More on inverse square orbits

Having got a good physical insight into the general nature of central force motion, the
last few sections focus on details of motion under inverse square forces. Remember that

q1 q2
F(r) = 2 r, where = GMm for Newtonian gravity (always attractive) and =
r
40
for the Coulomb interaction (which can be both attractive and repulsive).

5.1

Geometrical aspects

First some comments on geometrical aspects of the motion and how to fix the parameters.

5.1.1

Bounded Orbits

The orbit is an ellipse, examples are planets and comets moving around a star. Fig. 7(a)
shows an ellipse with the origin of polar coordinates at a focus, F , which is also the centre
of the force. Such orbits are only possible for an attractive force. The equation of the

ellipse in polar coordinates is = 1 + e cos where and e (0 < e < 1) are constants.
r
Some features of the orbit:
P point of closest approach to F , (perigee), = 0, r = /(1 + e);
A point furthest from F , (apogee), = , r = /(1 e).
At both A and P , the radial velocity r is zero.
7

Figure 7: Inverse square force orbits: (a) ellipse for bounded orbit; (b) hyperbola for
unbounded orbits.
5.1.2

Unbounded Orbits

The open or unbounded orbit is a hyperbola, as shown in Fig. 7(b). The polar equation
has the same form as above but e > 1. The centre of force is at the focus F and is
measured from the distance of closest approach. The hyperbola has two branches, only
one of which will be valid for a particular motion:
Attractive force. The near branch, point P on the trajectory.
Repulsive force. The far branch, point P on the trajectory.
Examples of unbounded orbits: very energetic comets passing the sun; Rutherford scattering of sub-atomic particles.
5.1.3

Determination of the orbit parameters

Knowing the velocity at a given point on the orbit fixes its parameters. Apogee and
perigee are good choices because r = 0 there and the velocity r is at right angles to the
radius vector.
Suppose a satellite is launched into its orbit at perigee ( = 0), a distance d from the
focus, with velocity v.

From the equation of the orbit = 1 + e.


d
md2 v 2
.
From = mh2 / where h = r 2 = dv, so =

5.1.4

Orbit parameters and the total energy

Starting from the energy conservation equation




h2

1
2
m r + 2 = E
2
r
r

(11)

and the equation of the orbit = 1 + e cos it requires a few lines of algebra [try it] to
r
show that
2
2mh2
2
2
E=
(e

1)
or
e
=
E + 1.
(12)
2mh2
2
Thus:
8

Open orbits, E > 0 e2 > 1, hyperbolic;


Closed orbits, E < 0 e2 < 1, elliptical;
Special case, E = 0 or e = 1, parabolic.
Another way of remembering the difference is that for open orbits KE > PE, for close
orbits KE < PE.
5.1.5

Keplers Third Law

dA
h

= , where h = r 2 .
dt
2
h
Thus in period T the complete area enclosed by the orbit, A, is swept out and A = T .
2
For a planet in a circular orbit about the Sun with radius a and angular speed this gives
GM
.
T = 2/. Equating the attractive central force to the centrifugal force gives 2 =
a3
Thus
(2)2
(2)2 a3
T2 =
=
which is Kepler s third law.
2
GM
Start from Keplers 2nd Law,

For the general elliptical case the area is given by ab, where 2a, 2b are the major and
minor axis lengths.1 Thus
T =

2ab
h

a = /(1 e2 ) and b2 = a2 (1 e2 ) (by comparing the polar and cartesian equations


for an ellipse) a is the semi-major axis length.
from = mh2 /, h2 = a(1 e2 )/m
so T 2 =

(2)2 a2 b2
(2)2 a4 (1 e2 )m
(2)2 a3
=
=
h2
a(1 e2 )
GM

where = GM m has been used in the final step. For the Solar System the data on
KIII is shown in Fig 8.
The above calculation has assumed essentially a fixed centre of force, which is a very
good approximation for the Solar System as the Sun contains 99.8% of the total mass.
For other systems, such as binary stars, one has to separate the motion into motion of
the system overall and motion within the CM. Keplers Laws applies to the latter. The
formulation of KIII then needs modification and becomes
T2 =

(2)2 a312
,
G(m1 + m2 )

where m1 , m2 are the masses of the two stars and a12 is the distance between them.
1

The equation of an ellipse in cartesian form is

x2
y2
+ 2 = 1.
2
a
b

Period (years)

Solar System orbits

10 2

10

10

-1

10

-1

10
Semi-major axis (AU)

Figure 8: Keplers third law for the Solar System. A double log plot showing period T
as a function of semi-major axis length for the planets of the Solar System. The line is
T = const.a3/2 . AU is the astronomical unit and is the semi-major axis length of the
Earths orbit.

5.2

Distance of closest approach

Clearly, if the equation of the orbit is known in polar form, one may calculate immediately
the distance of closest approach to the focus (perigee) using rmin = /(1 + e).
However it is often the case that other information is known, but not the orbit. Refering
to Fig. 9, a useful concept is the impact paramter b. A mass m a very long way away
from the force centre at F has velocity v a perpendicular distance b from a line parallel
1
to v through F . The initial energy of the mass is E = mv2 and the initial angular
2
momentum, or h parameter is h = bv. Rewrite the energy equation (11) in the form


E
h2

2
r = 2
2.
+
m mr
r
At the distance of closest approach, r = 0, which gives a quadratic equation for r

2
2 + 2Emh2

+
mh
= 0 with solution rmin =
.
Er 2 + r
2
2E

(13)

An immediate result is that the force centre cannot be reached (rmin = 0), unless the
initial impact parameter b = 0. Consider a topical example, suppose an asteroid is seen
approaching the Earth, will it collide? Is rmin < REarth ? This could be true even though
b 6= 0, hence the recent intergovernmental decision to set up a programme to find nearby
asteroids and map their orbits.

10

Figure 9: Impact parameter b and distance of closest approach for a mass m with velocity
v a long way away from the force centre F .

5.3

Angle of deflection; Rutherford scattering

This again concerns unbound orbits by what angle is the path of a particle or asteroid
turned asymptotically in its encounter with a force centre? Refer again to Fig. 9. The
angle required is labelled and is the angle between the two asymptotes to the hyperbolic
trajectory. Polar angles are measured from the position of closest approach at D, then
is the polar direction of the mass along its initial trajectory. [Note: the results in this
section apply to both repulsive and attractive forces. The diagrams are for an attractive
force.]
5.3.1

Geometrical approach

There are two approaches, the first uses = 1 + e cos . The position of closest approach
r
D lies on the bisector of the between the asymptotes, thus
= ( )/2 or = 2 and sin

= cos .
2

Initial = and r is very large, thus cos = 1/e. So


cot2

2Emh2

= e2 1 =
,
2
2

where Eqn 12 has been used. Initially E =


written cot
5.3.2

mbv 2

=
.
2

1 2
mv and h = bv, so the result may also be
2

Elastic Scattering

The disadvantage of the above method is that it assumes a detailed knowledge of the
properties of the hyperbolic trajectory in polar form. An alternative is to consider the
interaction as two-body elastic scattering. The initial momentum mv is not changed in
magnitude but its direction is changed. That change may be calculated from the integrated impulse provided by the force. Fig. 10 shows the mass m with initial velocity v at
impact parameter b scattered by the (attractive) force at F . The symmetry of the trajectory about = 0 line ADF will again be important. In addition to the polar coordinates,
11

Figure 10: Coordinate systems for calculating the angle of deflection by the impulse
method.
resolve v along two fixed directions: vk parallel to F DA; v perpendicular (shown as
the small inserts). From the symmetry about ADF it is clear that asymptotically v is
unchanged, but vk is turned around. The change in momentum is related to the integral
over the force
Z
Z 2
mr 2
d,
F
(mv) =
F dt =
J

where the conservation of AM J = mr 2 has been used to change the variable of inte
gration from time to angle. Now F = 2 r and at a point P (r, ) on the trajectory
r
r = (cos , sin ) when resolved into components (k, ) to F DA. Thus
Z
m 2
m(vk , v ) =
(cos , sin )
J
m

=
[sin , cos ]2

J
m
(2 sin , 0).
=
J
So mv = 0, in agreement with the symmetry argument and
2m
sin
J
= 2mv cos ,

|mvk | =

using the projection of v at initial and final positions. Finally




Jv

where J = mbv.
| tan | = cot =
2

12

A clever use of this result in the gravitational case, is the scattering of spacecraft off
the inner planets or even Jupiter to swing them further towards the outer solar system or
beyond, thus saving on launch energy.
5.3.3

Rutherford Scattering

In the classic experiments of Rutherford, Geiger and Marsden in 1911 the structure of the
nuclear atom was discovered. At the time there were no accelerators so particle beams for
scattering experiments came from collimating particles emitted from energetic radioactive
decays. In -decay an unstable nucleus emits a helium nucleus (a particle with roughly 4
times the proton mass and twice its charge). The helium nuclei ( particles) may then be
scattered from other nuclear targets here provided by a very thin gold foil. Rutherford
was suprised that sometimes the particles were scattered through very large angles
almost 180 . He realised that this could only happen if the total positive charge of the
atom was concentrated essentially at a point. Using classical mechanics and his nuclear
model, he was able to calculate what one should expect.
What is known is the initial -particle energy E and direction, and one measures the
distribution of scattering angles (what is denoted by in the above discussion). One
cannot measure the impact parameter of scattering events, however one may use the
cot(/2)
results of this section to replace b or h by cot(/2), b =
. Using this in Eqn 13
2E
gives
p
+ 2 + 2 cot2 (/2)
.
rmin =
2E
2Ze2
, where Z is
For the classic Rutherford scattering the force is repulsive and =
40
the charge on the target nucleus. The above formula becomes

 2 
q
Z
e
2
rmin =
1 + 1 + cot (/2)
E 40
which shows the dependence of rmin on Z, E and the scattering angle . Later experiments
by Rutherford showed that the nucleus, although very small on the atomic scale, did
nonetheless have a finite size. For large E and , rmin is less that the nuclear size
and deviations are observed from the expectations of electromagnetic Coulomb scattering.
For -nucelus scattering at these distances, the strong nuclear force comes into play.
The technique of scattering as a probe of sub-atomic structure continues to this day, but
using accelerators. The most accurate probes use electron or positron beams for which
Ze2
=
and the scattering is attractive (electrons) or repulsive (positrons).
40

5.4

CM frame

In all discussions of planetary motion and scattering so far, the force centre has been
assumed to be fixed. For planets orbitting the Sun and scattering off a heavy nucleus,
this is a good assumption. It is not so good for objects such as binary stars or even
the Earth-Moon system. The solution has already been discussed see Section 3 of the
notes on systems of point particles it is to refer the motion to the centre-of-mass of
the system. The forces are split into external forces acting on the system as a whole
and the internal forces. For the former the system behaves as a point particle with
13

total mass at the CM. In the case of a two-body system, the internal motion reduces to
an equivalent one-body problem in terms of the relative position vector and the reduced
mass = m1 m2 /(m1 + m2 ).
r = F12 where r = r1 r2
and F12 is the internal force between m1 and m2 . Thus all the above results may be
taken over with these two changes. As an example, consider the Earth-Moon system
orbitting the Sun (Fig. 11). The Earth and Moon revolve about their common CM and
the system revolves about the Sun. Because the Sun is so much more massive than the
combined system, one may approximate the motion about the Sun as that of a point mass
(m = MEarth + MM oon ) about a fixed centre of attraction. The resulting motion of the
moon in space is shown in the RH diagram of the figure. The wobble of the Earths
orbit is of course much smaller.

Figure 11: Motion of the Earth and Moon around each other and around the Sun.

14

You might also like